\jyear

2023

\equalcont

These authors contributed equally to this work.

\equalcont

These authors contributed equally to this work.

[4]\fnmVasili \surPerebeinos

[2,12]\fnmAndrey \surBaydin

[1,2,12,13]\fnmJunichiro \surKono

[1,2,12,13]\fnmHanyu \surZhu

1]\orgdivDepartment of Materials Science and NanoEngineering, \orgnameRice University, \orgaddress\cityHouston, \postcode77005, \stateTexas, \countryUSA

2]\orgdivDepartment of Electrical and Computer Engineering, \orgnameRice University, \orgaddress\cityHouston, \postcode77005, \stateTexas, \countryUSA

3]\orgdivApplied Physics Graduate Program, Smalley-Curl Institute, \orgnameRice University, \orgaddress\street \cityHouston, \postcode77005, \stateTexas, \countryUSA

4]\orgdivDepartment of Electrical Engineering, \orgnameUniversity at Buffalo, \orgaddress\cityBuffalo, \postcode14228, \stateNew York, \countryUSA

5]\orgnameJ.A. Woollam Co., Inc., \orgaddress\cityLincoln, \postcode68508, \stateNebraska, \countryUSA

6]\orgdivLaboratory of Biomaterials and Intelligent Materials, \orgnameLebanese University, \orgaddress\cityJdeidet, \postcode90656, \countryLebanon

7]\orgdivDepartment of Physics, \orgnameLebanese University, \orgaddress\cityJdeidet, \postcode90656, \countryLebanon

8]\orgdivDepartment of Electrical and Computer Engineering, \orgnameUniversity of Utah, \orgaddress\citySalt Lake City, \postcode84112, \stateUtah, \countryUSA

9]\orgdivDepartment of Physics, \orgnameTokyo Metropolitan University, \orgaddress\cityTokyo, \postcode192-0397, \countryJapan

10]\orgdivDepartment of Physics, \orgnameTohoku University, \orgaddress\citySendai, \postcode980-8578, \countryJapan

11]\orgdivDepartment of Physics, \orgnameNational Taiwan Normal University, \orgaddress\cityTaipei, \postcode106, \countryTaiwan

12]\orgdivSmalley-Curl Institute, \orgnameRice University, \orgaddress\cityHouston, \postcode77005, \stateTexas, \countryUSA

13]\orgdivDepartment of Physics and Astronomy, \orgnameRice University, \orgaddress \cityHouston, \postcode77005, \stateTexas, \countryUSA

Giant Second Harmonic Generation from Wafer-Scale Aligned Chiral Carbon Nanotubes

\fnmRui \surXu rx8@rice.edu    \fnmJacques \surDoumani jd74@rice.edu    Viktor Labuntsov labuntsov99@mail.ru    \fnmNina \surHong nhong@jawoollam.com    \fnmAnna-Christina \surSamaha annachristina.samaha@st.ul.edu.lb    \fnmWeiran \surTu wt20@rice.edu    \fnmFuyang \surTay ft13@rice.edu    \fnmElizabeth \surBlackert erb13@rice.edu    \fnmJiaming \surLuo jl230@rice.edu    \fnmMario \surEl Tahchi mtahchi@ul.edu.lb    \fnmWeilu \surGao weilu.gao@utah.edu    \fnmJun \surLou jlou@rice.edu    \fnmYohei \surYomogida yomogida@tmu.ac.jp    \fnmKazuhiro \surYanagi yanagi-kazuhiro@tmu.ac.jp    \fnmRiichiro \surSaito rsaito@tohoku.ac.jp    vasilipe@buffalo.edu    baydin@rice.edu    kono@rice.edu    hz67@rice.edu [ [ [ [ [ [ [ [ [ [ [ [ [
Abstract

Chiral carbon nanotubes (CNTs) are direct-gap semiconductors with optical properties governed by one-dimensional excitons with enormous oscillator strengths. Each species of chiral CNTs has an enantiomeric pair of left- and right-handed CNTs with nearly identical properties, but enantiomer-dependent phenomena can emerge, especially in nonlinear optical processes. Theoretical studies have predicted strong second-order nonlinearities for chiral CNTs, but there has been no experimental verification due to the lack of macroscopically ordered assemblies of single-enantiomer chiral CNTs. Here for the first time, we report the synthesis of centimeter-scale films of densely packed and aligned single-enantiomer chiral CNTs that exhibit micro-fabrication compatibility. We observe giant second harmonic generation (SHG) emission from the chiral CNT film, which originates from the intrinsic chirality and inversion symmetry breaking of the atomic structure of chiral CNTs. The observed value of the dominant element of the second-order nonlinear optical susceptibility tensor reaches 1.5 ×\times× 103 pm/V at a pump wavelength of 1030 nm, corresponding to the lowest-energy excitonic resonance. Our calculations based on many-body theory correctly estimate the spectrum and magnitude of such excitonically enhanced optical nonlinearity. These results are promising for developing scalable chiral-CNT electronics, nonlinear photonics and photonic quantum computing.

Main Text

Structural chirality is a fundamental property in which the material is not identical to its mirror image. Materials possessing chirality can enable intriguing properties, including optical activity noauthor_circular_nodate ; ahn_new_2017 , magneto-chiral effects rikken_observation_1997 ; train_strong_2008 , and chiral-induced spin selectivity naaman_chiral-induced_2012 ; yang_real-time_2023 , which can be useful for information processing and transfer applications yang_circularly_2013 ; brandt_added_2017 ; yang_chiral_2021 ; wei_chiral_2021 . Due to lifted inversion symmetry in chiral materials, second-order optical nonlinearity is also expected for conventional chiral crystalline insulators including α𝛼\alphaitalic_α-quartz and TeO2 singh_violation_1972 ; chemla_optical_2003 and recently discovered chiral semiconductors such as hybrid organic-inorganic halides yuan_chiral_2018 ; yao_strong_2021 ; ge_chiral_2022 , with resonance-enhanced nonlinear optical susceptibilities χ(2)superscript𝜒2\chi^{(2)}italic_χ start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT of up to 100 pm/V. However, it is counter-intuitive that second harmonic generation (SHG) can arise purely from structural chirality without any local electrical or magnetic polarization, such as in non-magnetic elemental allotropes. Indeed, Kleinman symmetry dictates that SHG is forbidden in isotropic chiral materials without dispersions boyd_nonlinear_1992 . Previously, crystalline anisotropy and large dispersions have enabled large SHG from tellurium, a chiral elemental allotrope cheng_large_2019 ; londono-calderon_intrinsic_2021 ; wang_anisotropic_2021 , but the nonlinear susceptibility was only quantified at far-infrared wavelengths, instead of near the resonance at the bandgap shermanAbsoluteMeasurementSecondharmonic1973 .

Chiral carbon nanotubes (CNTs) are ideal candidates for exploring chirality-induced nonlinear optical phenomena. Since their discovery, CNTs have shown unique electronic, photonic, and mechanical properties arising from one-dimensional (1D) quantum confinement effects AvourisEtAl2008NP ; BlackburnEtAl2018AMb ; Li2021AN ; BaydinEtAl2022M . Calculations have predicted strong 1D van Hove singularity-enhanced second-order nonlinearities, χ(2)superscript𝜒2\chi^{(2)}italic_χ start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT on the order of nm/V, for chiral species of single-wall CNTs, where their chirality indices (n𝑛nitalic_n,m𝑚mitalic_m) satisfy nm𝑛𝑚n\neq mitalic_n ≠ italic_m and m0𝑚0m\neq 0italic_m ≠ 0 guo_linear_2004 ; rerat_ab_2018 . Such chiral CNTs are semiconductors with diameter (dtsubscript𝑑td_{\rm t}italic_d start_POSTSUBSCRIPT roman_t end_POSTSUBSCRIPT)-dependent direct band gaps weisman_introduction_2017 , Egsubscript𝐸gE_{\rm g}italic_E start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT (eV) similar-to\sim 0.7/dtsubscript𝑑td_{\rm t}italic_d start_POSTSUBSCRIPT roman_t end_POSTSUBSCRIPT (nm), potentially useful for integrated nanophotonics and chiral optoelectronics. Being 1D semiconductors with dt1similar-tosubscript𝑑t1d_{\rm t}\sim 1italic_d start_POSTSUBSCRIPT roman_t end_POSTSUBSCRIPT ∼ 1 nm, chiral CNTs exhibit enormous excitonic effects, which can resonantly boost SHG processes further nishihara_empirical_2022 . Several experiments have observed SHG from different types of CNT materials, including single tubes huttunen_measurement_2013 , thin films okawara_second-harmonic_2019 ; dominicis_symmetry_2004 , and chiral CNTs embedded in host porous crystals su_resonant_2008 . However, the observed optical nonlinearity was very weak (effective χ(2)0.01similar-tosuperscript𝜒20.01\chi^{(2)}\sim 0.01italic_χ start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT ∼ 0.01 pm/V) due to the coexistence of both enantiomers, random CNT orientations, and low packaging densities, hindering both fundamental studies and practical applications. Furthermore, the observed SHG may have originated from surface effects or local defects instead of the intrinsic chirality of CNTs since the symmetry of the observed signal disagreed with the expected SHG tensor of chiral CNTs, which should contain only two non-zero elements, i.e., χxyz(2)subscriptsuperscript𝜒2𝑥𝑦𝑧\chi^{(2)}_{xyz}italic_χ start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_x italic_y italic_z end_POSTSUBSCRIPT and χyzx(2)subscriptsuperscript𝜒2𝑦𝑧𝑥\chi^{(2)}_{yzx}italic_χ start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_y italic_z italic_x end_POSTSUBSCRIPT (=χxyz(2)absentsubscriptsuperscript𝜒2𝑥𝑦𝑧=-\chi^{(2)}_{xyz}= - italic_χ start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_x italic_y italic_z end_POSTSUBSCRIPT).

Here, we report the fabrication of centimeter-scale films of aligned enantiomer-pure chiral CNTs and the observation of giant SHG from the films. We chose (6,5)- CNTs for this study; (6,5)+ and (6,5)- are the left-handed and right-handed enantiomers, respectively, of the otherwise identical CNT species with dt=0.76subscript𝑑t0.76d_{\rm t}=0.76italic_d start_POSTSUBSCRIPT roman_t end_POSTSUBSCRIPT = 0.76 nm and Eg1similar-tosubscript𝐸g1E_{\rm g}\sim 1italic_E start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT ∼ 1 eV. Incident angle- and polarization-dependent measurements of SHG, combined with analysis of the χ(2)superscript𝜒2\chi^{(2)}italic_χ start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT tensor, confirmed that the observed SHG signal originated from the intrinsic structural chirality of the (6,5)- CNTs. The excitation-wavelength-dependence of the SHG signal exhibited a prominent peak at around the E11subscript𝐸11E_{11}italic_E start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT exciton resonance (similar-to\sim1 μμ\upmuroman_μm), in agreement with the results of our calculations based on many-body theory. Excitation power dependence showed a deviation from the expected quadratic behavior for excitation intensities above a threshold of 0.2similar-toabsent0.2\sim 0.2∼ 0.2 mJ/cm2, which we attribute to absorption saturation due to exciton-exciton annihilation murakami_nonlinear_2009 . From the measured SHG signals, we estimated the value of χxyz(2)subscriptsuperscript𝜒2𝑥𝑦𝑧\chi^{(2)}_{xyz}italic_χ start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_x italic_y italic_z end_POSTSUBSCRIPT to be 1.5×1031.5superscript1031.5\times 10^{3}1.5 × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT pm/V, which is comparable with values reported for single-crystalline two-dimensional semiconductors and Weyl semimetals janisch_extraordinary_2014 ; kumar_second_2013 ; you_nonlinear_2019 ; kim_three-dimensional_2024 ; wu_giant_2017 . These results demonstrate that aligned enantiomer-pure CNT films, which can be easily transferred to any substrate with excellent conformity and scalability, are promising for applications in integrated chiral nano-photonics and chiral quantum optics lodahl_chiral_2017 .

We first prepared an aqueous suspension of enantiomer-enriched (6,5)- CNTs and measured its enantiomer purity (EP𝐸𝑃EPitalic_E italic_P). The suspension was purified using gel chromatography yomogida_industrial-scale_2016 ; yomogida_automatic_2020 ; wei_high-yield_2018 , as illustrated in Fig. 1a and described in more detail in Methods. The EP𝐸𝑃EPitalic_E italic_P of the suspension was characterized by circular dichroism (CD) spectroscopy; see Fig. 1b. The CD intensity spectrum exhibits multiple peaks corresponding to the Eijsubscript𝐸𝑖𝑗E_{ij}italic_E start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT excitonic transitions shown in the attenuation spectrum; here, i𝑖iitalic_i and j𝑗jitalic_j are integers, and transitions with i=j𝑖𝑗i=jitalic_i = italic_j are strongly allowed for light polarization parallel to the nanotube axis WeismanKono19Book . The EP𝐸𝑃EPitalic_E italic_P of the (6,5)- CNT suspension can be calculated from the CD intensity (in mdeg) normalized by the attenuance A𝐴Aitalic_A at the E22subscript𝐸22E_{22}italic_E start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT transition, denoted by CDE22𝐶subscript𝐷subscript𝐸22CD_{E_{22}}italic_C italic_D start_POSTSUBSCRIPT italic_E start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT end_POSTSUBSCRIPT and AE22subscript𝐴subscript𝐸22A_{E_{22}}italic_A start_POSTSUBSCRIPT italic_E start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT end_POSTSUBSCRIPT, respectively, as follows wei_determination_2017 :

EP(6,5)=(500.421CDE22AE22)%=91.6%,𝐸𝑃superscript65percent500.421𝐶subscript𝐷subscript𝐸22subscript𝐴subscript𝐸22percent91.6\displaystyle EP(6,5)^{-}=\left(50-0.421\frac{CD_{E_{22}}}{A_{E_{22}}}\right)% \%=91.6\%,italic_E italic_P ( 6 , 5 ) start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT = ( 50 - 0.421 divide start_ARG italic_C italic_D start_POSTSUBSCRIPT italic_E start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_ARG start_ARG italic_A start_POSTSUBSCRIPT italic_E start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_ARG ) % = 91.6 % , (1)

where we used CDE22=29.67𝐶subscript𝐷subscript𝐸2229.67CD_{E_{22}}=-29.67italic_C italic_D start_POSTSUBSCRIPT italic_E start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT end_POSTSUBSCRIPT = - 29.67 mdeg and AE22=0.30subscript𝐴subscript𝐸220.30A_{E_{22}}=0.30italic_A start_POSTSUBSCRIPT italic_E start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT end_POSTSUBSCRIPT = 0.30; see Supporting Information Section 1 for zoom-in data. This EP𝐸𝑃EPitalic_E italic_P value is comparable with the state-of-art in the literature wei_length-dependent_2023 .

Refer to caption
Figure 1: Fabrication and characterization of an aligned film of enantiomer-pure (6,5)- CNTs. a, An as-purchased mixture of CNTs was dispersed in water and separated by their metallicity and chirality via multi-step gel chromatography. A (6,5)- CNT suspension was extracted, and an aligned film was made by using controlled vacuum filtration (CVF); the fabricated film was transferred onto a fused silica substrate. b, Circular dichroism and attenuance spectra for the (6,5)- CNT suspension. Analysis of these spectra yielded an enantiomer purity of 91.6%. c, An optical image of the chiral aligned (CA) CNT thin film with a diameter of 3 cm. d,e, SEM images of CA CNT thin film at low and high magnifications, respectively, demonstrating the film’s uniformity, high packing density, and microscopic in-plane anisotropy. The yellow arrow indicates the alignment direction. f, Global linearly polarized attenuance spectra for the CA sample with a probed area of about 1.2×1031.2superscript1031.2\times 10^{3}1.2 × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPTμμ\upmuroman_μm2. The observed peaks are due to the E11subscript𝐸11E_{11}italic_E start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT exciton resonance, the E11subscript𝐸11E_{11}italic_E start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT exciton phonon sideband (PSB), and the E22subscript𝐸22E_{22}italic_E start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT exciton resonance. The linear dichroism value at E11subscript𝐸11E_{11}italic_E start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT is 0.9, corresponding to a nematic order parameter of similar-to\sim0.45.

We then fabricated three CNT films using the CVF method he_wafer-scale_2016 and transferred them onto fused silica substrates (Fig. 1c and Methods). The first film, labeled ‘chiral aligned’ or CA, contained aligned enantiomer-pure (6,5)- CNTs. The film preserved the single-chirality nature of the original CNT suspension, as evidenced by Raman spectra (Supporting Information Section 1) PhysRevB.72.205438 . The second sample, labeled ‘chiral random’ or CR, contained randomly oriented enantiomer-pure (6,5)- CNTs prepared from the same suspension as the CA film. Finally, the third film, labeled ‘racemic aligned’ or RA, consisted of highly aligned CNTs containing metallic and semiconducting CNTs with mixed chiralities he_wafer-scale_2016 . All three films had the same thickness of around 20 nm, measured by atomic force microscopy (Supporting Information Section 2). Scanning electron microscope images indicated that these films were densely packed with CNTs with no visible voids at optical wavelength scales (Fig. 1d and Supporting information Section 3).

Polarized attenuance spectra for the CA film (Fig. 1f) confirmed its global in-plane anisotropy within a probed area of 1.2×1031.2superscript1031.2\times 10^{3}1.2 × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPTμμ\upmuroman_μm2. The first three peaks in the attenuance spectra correspond to the E11subscript𝐸11E_{11}italic_E start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT exciton resonance (1.23 eV), the phonon sideband of the E11subscript𝐸11E_{11}italic_E start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT exciton resonance (PSB, 1.43 eV), and the E22subscript𝐸22E_{22}italic_E start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT exciton resonance (2.15 eV), respectively. The difference in E11subscript𝐸11E_{11}italic_E start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT attenuance between parallel (AHsubscript𝐴𝐻A_{H}italic_A start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT) and perpendicular (AVsubscript𝐴𝑉A_{V}italic_A start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT) polarizations with respect to the CNT alignment direction yields a linear dichroism (LD) of 2×(AHAV)/(AH+AV)=0.92subscript𝐴Hsubscript𝐴Vsubscript𝐴Hsubscript𝐴V0.92\times(A_{\rm H}-A_{\rm V})/(A_{\rm H}+A_{\rm V})=0.92 × ( italic_A start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT - italic_A start_POSTSUBSCRIPT roman_V end_POSTSUBSCRIPT ) / ( italic_A start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT + italic_A start_POSTSUBSCRIPT roman_V end_POSTSUBSCRIPT ) = 0.9, corresponding to a nematic order parameter S=0.45𝑆0.45S=0.45italic_S = 0.45 within the probed area katsutani_direct_2019 . On a more local scale within an optically resolvable area of 1.3 μμ\upmuroman_μm2, the CA film showed stronger alignment with S=0.660.71𝑆0.66similar-to0.71S=0.66\sim~{}0.71italic_S = 0.66 ∼ 0.71 (Supporting Information Section 4). The random orientation of CNTs in the CR film was also confirmed (S=0.04𝑆0.04S=0.04italic_S = 0.04) by its lack of LD (Supporting Information Section 5) as well as SEM images (Supporting Information Section 3).

Refer to caption
Figure 2: Quantifying the SHG tensor elements of aligned enantiomer-pure (6,5)- CNT thin films. a, The excitation pulse is horizontally polarized (z𝑧zitalic_z-axis), whose electric field is at an angle θ𝜃\thetaitalic_θ with respect to the CNT alignment direction. The emitted SHG is separated from the excitation pulse by an optical filter and analyzed via a polarizer at an angle φ𝜑\varphiitalic_φ from the excitation polarization. Axes with (without) the prime denote the lab (sample) coordinate system. Inset shows the chiral atomic structure of a single (6,5)-, a left-handed CNT. b, Strong SHG emission is observed in the chiral enantiomer-pure and aligned (CA) film under non-zero tilting angle, as expected from the χyxz(2)superscriptsubscript𝜒𝑦𝑥𝑧2\chi_{yxz}^{(2)}italic_χ start_POSTSUBSCRIPT italic_y italic_x italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT element, compared with negligible SHG from either the chiral random (CR) or racemic aligned (RA) films. c, The emitted SHG radiation is vertically polarized, which is orthogonal to the excitation pulses, again in agreement with the symmetry of the SHG tensor. Here, the fitting parameter φ0subscript𝜑0\varphi_{0}italic_φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the angle between the SHG polarization and the z𝑧zitalic_z axis in the laboratory system.

Next, we resonantly pumped the CA film in the vicinity of E11subscript𝐸11E_{11}italic_E start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT exciton transition energy (1.2similar-toabsent1.2\sim 1.2∼ 1.2 eV) and measured emitted radiation at the second harmonic (2.4similar-toabsent2.4\sim 2.4∼ 2.4 eV). In chiral CNTs, such as (6,5) CNTs, only one independent nonzero element, χxyz(2)=χyxz(2)superscriptsubscript𝜒superscript𝑥superscript𝑦superscript𝑧2superscriptsubscript𝜒superscript𝑦superscript𝑥superscript𝑧2\chi_{x^{\prime}y^{\prime}z^{\prime}}^{(2)}=-\chi_{y^{\prime}x^{\prime}z^{% \prime}}^{(2)}italic_χ start_POSTSUBSCRIPT italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_y start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT = - italic_χ start_POSTSUBSCRIPT italic_y start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT, exists in the second order nonlinear optical susceptibility tensor guo_linear_2004 , where zsuperscript𝑧z^{\prime}italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT denotes the nanotube axis in the sample coordinate system (Fig. 2a and Supporting Information Section 7). Therefore, the intensity of SHG, ISHGsubscript𝐼SHGI_{\rm SHG}italic_I start_POSTSUBSCRIPT roman_SHG end_POSTSUBSCRIPT, arising from the intrinsic structural chirality should be nonzero only when the electric field of the incident light has finite projections onto both the ysuperscript𝑦y^{\prime}italic_y start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT and zsuperscript𝑧z^{\prime}italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT axes in the sample coordinate system:

ISHGPx2=(2|χxyz(2)|SEyEz)2,proportional-tosubscript𝐼SHGsuperscriptsubscript𝑃superscript𝑥2superscript2superscriptsubscript𝜒superscript𝑥superscript𝑦superscript𝑧2𝑆subscript𝐸superscript𝑦subscript𝐸superscript𝑧2\displaystyle I_{\rm SHG}\propto P_{x^{\prime}}^{2}=\left(2\lvert\chi_{x^{% \prime}y^{\prime}z^{\prime}}^{(2)}\rvert SE_{y^{\prime}}E_{z^{\prime}}\right)^% {2},italic_I start_POSTSUBSCRIPT roman_SHG end_POSTSUBSCRIPT ∝ italic_P start_POSTSUBSCRIPT italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = ( 2 | italic_χ start_POSTSUBSCRIPT italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_y start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT | italic_S italic_E start_POSTSUBSCRIPT italic_y start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT italic_E start_POSTSUBSCRIPT italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (2)

where S𝑆Sitalic_S is the nematic order parameter of the film, Eysubscript𝐸superscript𝑦E_{y^{\prime}}italic_E start_POSTSUBSCRIPT italic_y start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT and Ezsubscript𝐸superscript𝑧E_{z^{\prime}}italic_E start_POSTSUBSCRIPT italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT are the electric field projections of the fundamental field onto the ysuperscript𝑦y^{\prime}italic_y start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT and zsuperscript𝑧z^{\prime}italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT axes, respectively. The emission dipole Pxsubscript𝑃superscript𝑥P_{x^{\prime}}italic_P start_POSTSUBSCRIPT italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT is expected to be linearly polarized along the xsuperscript𝑥x^{\prime}italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT axis (Supporting Information Section 7). As illustrated in Fig. 2a, the CNT film was initially aligned with the polarization of the fundamental beam along the z𝑧zitalic_z-axis in the laboratory coordinate system and then tilted around the x𝑥xitalic_x-axis by an angle θ𝜃\thetaitalic_θ (Methods and Supporting Information Section 12). We observed strong SHG emission only from the CA film only when θ0𝜃superscript0\theta\neq 0^{\circ}italic_θ ≠ 0 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT; under normal incidence (θ=0𝜃superscript0\theta=0^{\circ}italic_θ = 0 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT), no SHG was observed due to the absence of Eysubscript𝐸superscript𝑦E_{y^{\prime}}italic_E start_POSTSUBSCRIPT italic_y start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT, as expected from Eq. 2. The angular dependence of the SHG emission intensity can be fit well by Eq. S17, which includes the effects of refraction and phase matching. At θ=45𝜃superscript45\theta=45^{\circ}italic_θ = 45 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT, we further confirmed that the polarization of the SHG radiation can be fit by cos2(φφ0)superscript2𝜑subscript𝜑0\cos^{2}(\varphi-\varphi_{0})roman_cos start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_φ - italic_φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ), where φ0subscript𝜑0\varphi_{0}italic_φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the angle between the SHG polarization and the z𝑧zitalic_z axis in the laboratory system. As shown in Fig. 2c, we used φ0subscript𝜑0\varphi_{0}italic_φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT as a fitting parameter and obtained φ0=87.5subscript𝜑0superscript87.5\varphi_{0}=87.5^{\circ}italic_φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 87.5 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT as the optimum value, indicating that the SHG emission is vertically polarized along the xsuperscript𝑥x^{\prime}italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT (x𝑥xitalic_x) axis in the sample (laboratory) coordinate system, consistent with Eq. 2. The combined tilting angle- and polarization-dependent measurements provided strong evidence that the observed SHG signal stems from the intrinsic chirality of the CNTs instead of surface symmetry breaking, which should yield ysuperscript𝑦y^{\prime}italic_y start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT-polarized emission, or local defects in the film, which should give a finite SHG signal under normal incidence. We observed small nonzero SHG from the CR film with an intensity less than 3% of that from the CA film, likely due to a finite nematic order through local spontaneous alignment. We observed no SHG emission from the RA film within the detection limit. These results highlight the essential role of both enantiomer purity and alignment for observing strong SHG emission from films of chiral CNTs.

Refer to caption
Figure 3: Photon-energy and fluence dependence of the observed second-order optical nonlinearity of the aligned enantiomer-pure (6,5)- CNT film. a, Experimentally determined |χxyz(2)|superscriptsubscript𝜒𝑥𝑦𝑧2\lvert\chi_{xyz}^{(2)}\rvert| italic_χ start_POSTSUBSCRIPT italic_x italic_y italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT | with a peak value approaching 1500  pm/V for excitation photon energy at around 1.2 eV near the E11subscript𝐸11E_{11}italic_E start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT exciton resonance (blue shade) and an additional peak at around 1 eV near half of the E12subscript𝐸12E_{12}italic_E start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT exciton resonance (purple shade), after correcting for the effects of multiple reflections, phase mismatch, CNT misalignment, and the collection efficiency of the setup. The width of the red-shaded area corresponds to the error range. b, The excitation fluence dependence of the SHG intensity (left axis) measured at a fundamental photon energy of 1.23 eV. The data shows a deviation from a quadratic fit above the saturation absorption threshold of similar-to\sim0.2  mJ/cm2 determined by power-dependent optical transmission measurements (right axis). The black arrow marks the excitation fluence used for taking the data shown in Fig. 3a, below which the SHG intensity exhibits a power-law dependence on the fluence with an optimum exponent of 2.04±0.06plus-or-minus2.040.062.04\pm 0.062.04 ± 0.06.

We determined the value of |χxyz(2)|superscriptsubscript𝜒𝑥𝑦𝑧2\lvert\chi_{xyz}^{(2)}\rvert| italic_χ start_POSTSUBSCRIPT italic_x italic_y italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT | from the SHG data for the CA film as a function of excitation photon energy in the range between 0.87 eV and 1.37 eV, finding strong excitonic enhancement (Figure 3a). In quantifying |χxyz(2)|superscriptsubscript𝜒𝑥𝑦𝑧2\lvert\chi_{xyz}^{(2)}\rvert| italic_χ start_POSTSUBSCRIPT italic_x italic_y italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT |, we considered multiple reflections, phase mismatch, CNT misalignment, and the collection efficiency of the setup (Supporting Information Section 7). The spectrum shows a prominent peak centered at 1.21 eV, closely matching the absorption resonance of the E11subscript𝐸11E_{11}italic_E start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT exciton observed in the linear optical susceptibility |χz(1)|superscriptsubscript𝜒𝑧1\lvert\chi_{z}^{(1)}\rvert| italic_χ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT | obtained through ellipsometry measurements (Supporting Information Section 8). The additional peak centered at 0.96 eV is attributed to the emission resonance with E12subscript𝐸12E_{12}italic_E start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT exciton (Figure 1b), whose optical dipole is in the xy𝑥𝑦xyitalic_x italic_y plane instead (Supplementary Information Section 9). The maximum measured χxyz(2)superscriptsubscript𝜒𝑥𝑦𝑧2\chi_{xyz}^{(2)}italic_χ start_POSTSUBSCRIPT italic_x italic_y italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT is 1.5×1031.5superscript1031.5\times 10^{3}1.5 × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT pm/V, which is the largest experimentally verified value in any 1D material systems, to the best of our knowledge, and is comparable with the highest χ(2)superscript𝜒2\chi^{(2)}italic_χ start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT values in the near-infrared region reported for any other nonlinear optical materials qian_large_2019 ; fu_berry_2023 ; zhang_large_2024 .

We further performed excitation fluence dependent measurements. The intensity of SHG emission from the CA film (red curve, left axis in Fig. 3b) increased quadratically as a function of excitation fluence Fpumpsubscript𝐹pumpF_{\rm pump}italic_F start_POSTSUBSCRIPT roman_pump end_POSTSUBSCRIPT. In the small-fluence region, with Fpumpsubscript𝐹pumpF_{\rm pump}italic_F start_POSTSUBSCRIPT roman_pump end_POSTSUBSCRIPT values up to 0.12 mJ/cm2, the data exhibited a power-law dependence with an optimum exponent k=2.04±0.06𝑘plus-or-minus2.040.06k=2.04\pm 0.06italic_k = 2.04 ± 0.06 when fit by the function ISHG=AFpumpksubscript𝐼SHG𝐴superscriptsubscript𝐹pump𝑘I_{\rm SHG}=AF_{\rm pump}^{k}italic_I start_POSTSUBSCRIPT roman_SHG end_POSTSUBSCRIPT = italic_A italic_F start_POSTSUBSCRIPT roman_pump end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT, as expected for SHG. However, as Fpumpsubscript𝐹pumpF_{\rm pump}italic_F start_POSTSUBSCRIPT roman_pump end_POSTSUBSCRIPT was further increased, a deviation from the quadratic dependence was observed. The deviation started at Fpump0.2similar-tosubscript𝐹pump0.2F_{\rm pump}\sim 0.2italic_F start_POSTSUBSCRIPT roman_pump end_POSTSUBSCRIPT ∼ 0.2 mJ/cm2, which coincided with the onset of saturation absorption measured by power-dependent transmission (blue curve, right axis in Fig. 3b). Such saturation behavior is commonly seen in CNT assemblies hussain_comparison_2019 , but the observed threshold fluence for the CA film was larger by one order of magnitude than what was previously reported in randomly oriented semiconducting CNT films containing multiple chiralities wang_152_2016 . This demonstrates that an aligned, densely packed, and enantiomer-pure CNT film is a more robust optical material. The data shown in Figures 2 and 3a were acquired by keeping Fpumpsubscript𝐹pumpF_{\rm pump}italic_F start_POSTSUBSCRIPT roman_pump end_POSTSUBSCRIPT below 0.12 mJ/cm2 (marked by a black arrow in Fig. 3b) to stay in the quadratic regime.

Refer to caption
Figure 4: Many-body atomistic calculations reproducing observed excitonic enhancement in (6,5)- CNTs at E11/E12subscript𝐸11subscript𝐸12E_{11}/E_{12}italic_E start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT / italic_E start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT resonances (blue/purple shade). (a) Linear absorption cross section for light polarized parallel (σzsubscript𝜎𝑧\sigma_{z}italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT) and perpendicular (σx,ysubscript𝜎𝑥𝑦\sigma_{x,y}italic_σ start_POSTSUBSCRIPT italic_x , italic_y end_POSTSUBSCRIPT) to the nanotube axis. (b) Nonlinear |χxyz(2)|superscriptsubscript𝜒𝑥𝑦𝑧2\lvert\chi_{xyz}^{(2)}\rvert| italic_χ start_POSTSUBSCRIPT italic_x italic_y italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT | spectrum exhibits large enhancement over the single-particle value when using the same damping rate. The features around 0.9 eV and 1.4 eV in the single-particle calculation were not seen in the experiment, further proving the importance of many-body effects in nonlinear optical processes in CNTs.

Finally, we offer a comprehensive theoretical understanding of the observed excitonic enhancement of optical nonlinearity via many-body calculations (Supporting Information Section 9). We first calculated the linear absorption cross-section of (6,5)- CNTs via second-order perturbation theory according to Ref. JiUng2013 , with the excitonic transition dipole moments evaluated using the Bethe–Salpeter equation Rohlfing2000 . As shown in Fig. 4a, symmetric excitonic absorption peaks are well-defined for light polarized parallel to the nanotube axis (σzsubscript𝜎𝑧\sigma_{z}italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT), which is qualitatively different from previous theoretical calculations for CNT systems based on single-particle band structure (Supporting Information Section 10). Introducing a Gaussian broadening parameter γ(E)=0.07eV+0.08eV(EE11)𝛾𝐸0.07eV0.08eV𝐸subscript𝐸11\gamma(E)=0.07\,\text{eV}+0.08\,\text{eV}\cdot(E-E_{11})italic_γ ( italic_E ) = 0.07 eV + 0.08 eV ⋅ ( italic_E - italic_E start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT ), which is proportional to the energy and loss channels (Supporting Information Section 9) Hertel2008 , reproduces the full-width-at-half-maximum (FWHM) of the E11subscript𝐸11E_{11}italic_E start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT and E22subscript𝐸22E_{22}italic_E start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT excitonic peaks in the polarized attenuance measurements (Fig. 1f). Then, following Ref. Boyd2020 , we calculated |χxyz(2)|superscriptsubscript𝜒𝑥𝑦𝑧2\lvert\chi_{xyz}^{(2)}\rvert| italic_χ start_POSTSUBSCRIPT italic_x italic_y italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT | spectra with the given γ𝛾\gammaitalic_γ and compared the results based on the excitonic picture with those based on the single-particle picture (Fig. 4b and Supporting Information Section 9).

Without any free parameters, the calculated peak positions and values of |χxyz(2)|superscriptsubscript𝜒𝑥𝑦𝑧2\lvert\chi_{xyz}^{(2)}\rvert| italic_χ start_POSTSUBSCRIPT italic_x italic_y italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT | from the excitonic picture agree with the experimental results well, in contrast to those from the single-particle calculations. The latter contains an artefact peak around 1.4 eV due to the contribution from higher-lying bands, which is a clear feature in all previous calculations but disagrees with the experiment. Another artifact around 0.9éV is due to the incorrect energy of E12=E21=(E11+E22)/2subscript𝐸12subscript𝐸21subscript𝐸11subscript𝐸222E_{12}=E_{21}=(E_{11}+E_{22})/2italic_E start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT = italic_E start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT = ( italic_E start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT + italic_E start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT ) / 2 whose degeneracy is lifted in the many-body picture by the exchange interaction. Previous reported |χxyz(2)|superscriptsubscript𝜒𝑥𝑦𝑧2\lvert\chi_{xyz}^{(2)}\rvert| italic_χ start_POSTSUBSCRIPT italic_x italic_y italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT | values based on the single-particle picture accidentally had the right order of magnitude by using an ideal constant γ=0.05𝛾0.05\gamma=0.05italic_γ = 0.05 eV for single CNT guo_linear_2004 ; pedersen_systematic_2009 , as opposed to our calculation using realistic and experimentally obtained γ𝛾\gammaitalic_γ values for densely packed thin films (Supporting Information Section 11). Meanwhile, we note that the many-body calculation appears to overestimate the nonlinearity at E12subscript𝐸12E_{12}italic_E start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT and E13subscript𝐸13E_{13}italic_E start_POSTSUBSCRIPT 13 end_POSTSUBSCRIPT resonance (the feature around 1.3éV). The fact that we did not observe a larger feature at 1.3éV implies that the damping rate of the E13subscript𝐸13E_{13}italic_E start_POSTSUBSCRIPT 13 end_POSTSUBSCRIPT transition in CNT films is larger than what we estimated in the simple model of linear energy dependence. Indeed, the E13subscript𝐸13E_{13}italic_E start_POSTSUBSCRIPT 13 end_POSTSUBSCRIPT resonance was not distinguishable in the linear absorption spectrum in experiments, making it impossible to evaluate its actual γ𝛾\gammaitalic_γ.

In summary, we discovered giant second-order optical nonlinearity in wafer-scale, enantiomer-pure, aligned, and densely packed chiral (6,5)- CNT thin films. The dependence of the SHG emission intensity on the sample orientation and optical polarization proves that the nonlinearity originates from intrinsic structural chirality instead of extrinsic contributions such as defects and surfaces. Control experiments with randomly oriented and mixed enantiomer thin films show that alignment and enantiomer purity are crucial for the observed chirality-induced SHG. Nonlinear spectroscopy reveals two excitonic resonances at pump photon energy of 1.21 eV and 0.96 eV, which corresponds to the absorption resonance with E11subscript𝐸11E_{11}italic_E start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT exciton and emission resonance with E12subscript𝐸12E_{12}italic_E start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT exciton, respectively. The second-order nonlinear susceptibility tensor element χxyz(2)superscriptsubscript𝜒𝑥𝑦𝑧2\chi_{xyz}^{(2)}italic_χ start_POSTSUBSCRIPT italic_x italic_y italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT reaches 1.5×103similar-toabsent1.5superscript103\sim 1.5\times 10^{3}∼ 1.5 × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT pm/V, comparable to the highest values experimentally reported in the near-infrared, in quantitative agreement with our many-body atomistic calculations with no free parameters. Our discovery opens a route for applying wafer-scale films of chiral CNTs to nonlinear and chiral photonics. Using our synthesis method, we can fabricate wafer-scale, flexible, and micro-fabrication-compatible thin films of chiral CNTs with various enriched diameters. Possibly allowing for broadband SHG response tunability across the spectrum, from the UV to NIR range.The strong nonlinearity of the film makes it potentially a good platform for mass scale quantum circuits components guo_ultrathin_2023 . Furthermore, since excitons in CNT are sensitive to external stimuli, including electrical gating chernikov_electrical_2015 and dielectric screening lin_dielectric_2014 , nonlinear chiral CNTs are tunable for possible optoelectronics and remote sensing.

Methods

Preparation of Enantiomer-Pure Single-Chirality (6,5)- CNTs

We used enantiomer-pure single-chirality (6,5)- CNTs (also known as (5,6) or (11,--5) or left-handed (6,5) CNTs), using the convention of Ref. wei_determination_2017 ; note that Refs. ghosh_advanced_2010 ; ao_differentiating_2016 used the opposite sign convention. In general, (n𝑛nitalic_n,m𝑚mitalic_m) CNTs are right-handed (left-handed) when n>m𝑛𝑚n>mitalic_n > italic_m (n<m)n<m)italic_n < italic_m ), and (n+m𝑛𝑚n+mitalic_n + italic_m,m𝑚-m- italic_m) CNTs are the mirror image of (n𝑛nitalic_n,m𝑚mitalic_m) CNTs PhysRevB.69.205402 . They were obtained by using gel chromatography employing mixed surfactant systems yomogida_industrial-scale_2016 ; yomogida_automatic_2020 ; wei_high-yield_2018 . CoMoCAT SWCNTs (704148, Sigma-Aldrich) were dispersed in 1.0% aqueous solution of sodium cholate (SC, >>>99%, Sigma-Aldrich) using an ultrasonicator (Sonifier 250D, Branson) with 30% output for three hours. After ultracentrifugation at 210,000g for two hours, the top 90% of the supernatant solution was collected and mixed with an aqueous solution of sodium dodecyl sulfate (SDS, >>>99.0%, Sigma-Aldrich) to prepare 2.0% SDS + 0.5% SC solution.

A two-step separation process using different surfactant concentrations yomogida_industrial-scale_2016 was performed automatically at 19-20 C using a high-performance liquid chromatography system (NGC chromatography system, BIO-RAD) with columns filled with gel beads (Sephacryl S-200 HR, GE Healthcare). First, the SWCNT solution was loaded into the first column equilibrated with 2.0% SDS + 0.5% SC solution, and non-adsorbed CNTs containing (6,5) CNTs were collected as filtrate. The filtrate solution is mixed with SC solution to prepare 0.5% SDS + 0.5% SC solution and used for the subsequent separation process. After loading the CNT solution into the second column equilibrated with 0.5% SDS + 0.5% SC solution, adsorbed CNTs were eluted by using mixed surfactant solution containing sodium lithocholate (LC, >>>98%, Tokyo Chemical Industry Co., Ltd.), i.e., 0.5% SDS + 0.5% SC + x𝑥xitalic_x% LC solution, where x𝑥xitalic_x is the LC concentration increased stepwise yomogida_automatic_2020 .

The high-purity (6,5)- CNTs were collected by adding a gradient solution of 0.5%SDS +0.5%SC+0.029% LC solution to 0.5%SDS +0.5%SC+0.030% LC solution. For film preparation, the contained surfactants were replaced with 0.3% sodium deoxycholate (DOC) by ultrafiltration (Amicon Ultra-15 with Ultracel-100 membrane, Merck Millipore). The obtained SWCNT solution was treated with ultrasonication with 30% output for ten minutes and ultracentrifugation at 210,000g for one hour to remove aggregates and used for film preparation (CA and CR films).

CNT Thin Film Preparation

For fabricating a CA film, the enantiomer-pure (6,5)- CNTs were aligned through the controlled vacuum filtration (CVF) method he_wafer-scale_2016 . The mother sodium deoxycholate (DOC) suspended CNT solution was diluted apriory to filtration by pure water to reduce the concentration of DOC from 0.3% to 0.075%. A total volume of 4 mL was filtered out through the filter membrane (Whatman Nuclepore Track-Etched membranes, MilliporeSigma). The filtration occurs in three stages: the gravitational, pumping, and drying phases to maintain a high controllability over the nematic ordering of the nanotubes. After filtration, the film was transferred onto a fused silica substrate (MTI corporation) by dissolving the polycarbonate membrane with chloroform, cleaning it with isopropanol (IPA), and blowing it with dry air. For fabricating a CR film, instead of diluting the solution with pure water, 1.5 mL of CNTs suspended in DOC with 0.3% concentration was added into 8.5 mL of diluted NaSCN (10 mmol/L of Na+). The suspension was then filtered out at a faster pace. The fast filtration and introduction of Na+ into the solution help make the alignment more chaotic. For fabricating a RA film, the whole recipe can be found in our previous work he_wafer-scale_2016 .

Characterization of Suspension and Aligned CNT Films

Sample morphology and thickness were characterized using a scanning electron microscope (FEI APREO) and an atomic force microscope (Park Systems NX20). Enantiometric purity (EP) was evaluated by recording the CD and attenuation spectra of the suspension using a CD spectrometer (Jasco 815). The uniaxially anisotropic refractive index was extracted from the Mueller matrix spectroscopic ellipsometry data acquired at three incident angles and three azimuthal orientations using the RC2 ellipsometer setup (JA Woollam Co.Inc).

Polarized Attenuance Spectra

Polarized attenuance spectra within a μμ\upmuroman_μm2 scale optically resolvable area for three samples were acquired by a home-built optical transmission measurement setup. A condensed lens was used to focus the light beam from the lamp onto the sample. The reference (transmitted) signal was collected by the objective under illumination at the area without (with) the sample. A polarizer was placed between the condenser lens and the sample to control the incident light polarization. The signal was then guided into the spectrometer through an optical fiber to get a spectrum. After subtracting noises from the spectrometer, the transmission was acquired by calculating the ratio of the transmitted spectrum over the reference spectrum. Assuming negligible reflection, attenuation was acquired by log(T)𝑇-\log(T)- roman_log ( italic_T ), where T𝑇Titalic_T is the intensity transmittance. By adjusting the magnification of the objective and the optical fiber’s diameter, the probing area size can be controlled.

Polarized Raman Spectroscopy

To comprehensively analyze nematic order, we employ a state-of-the-art Raman spectroscopy technique. Operating at a pump wavelength of 532 nm, we utilized a high-resolution Raman microscope equipped with a 50×50\times50 × objective lens, delivering a focused beam with a spot size of 1.3  μμ\upmuroman_μm2. To discern the alignment of nanotubes within the nematic medium, we investigated three distinct polarized configurations, VV, VH, and HH, while probing the same spot in all three configurations. The VV configuration aligned the incident and scattered light polarizations parallel to the nanotube axis, effectively probing alignment parallel to the incident beam. In the VH configuration, the incident polarization remained parallel to the nanotube axis while the scattered polarization becomes perpendicular, allowing us to gauge alignment perpendicular to the incident beam. Finally, the HH configuration employed perpendicular polarizations for both incident and scattered light. Additionally, by incorporating polarized attenuance data at 532 nm, using Equation LABEL:raman_eq, we assessed the nematic order parameter in three dimensions, which is a reasonable approximation since the thickness of the film is much larger than the diameter of the CNTs. With Δ=A/AΔsubscript𝐴parallel-tosubscript𝐴perpendicular-to\Delta=A_{\parallel}/A_{\perp}roman_Δ = italic_A start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT / italic_A start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT being the dichroic ratio raman , the nematic order parameter can be obtained as

SRaman=6ΔIVV+3(1+Δ)IVH8IHH6ΔIVV+12(1+Δ)IVH+16IHHsubscript𝑆Raman6Δsubscript𝐼VV31Δsubscript𝐼VH8subscript𝐼HH6Δsubscript𝐼VV121Δsubscript𝐼VH16subscript𝐼HH\displaystyle\begin{split}S_{\rm Raman}=\frac{6\Delta I_{\rm VV}+3(1+\Delta)I_% {\rm VH}-8I_{\rm HH}}{6\Delta I_{\rm VV}+12(1+\Delta)I_{\rm VH}+16I_{\rm HH}}% \\ \end{split}start_ROW start_CELL italic_S start_POSTSUBSCRIPT roman_Raman end_POSTSUBSCRIPT = divide start_ARG 6 roman_Δ italic_I start_POSTSUBSCRIPT roman_VV end_POSTSUBSCRIPT + 3 ( 1 + roman_Δ ) italic_I start_POSTSUBSCRIPT roman_VH end_POSTSUBSCRIPT - 8 italic_I start_POSTSUBSCRIPT roman_HH end_POSTSUBSCRIPT end_ARG start_ARG 6 roman_Δ italic_I start_POSTSUBSCRIPT roman_VV end_POSTSUBSCRIPT + 12 ( 1 + roman_Δ ) italic_I start_POSTSUBSCRIPT roman_VH end_POSTSUBSCRIPT + 16 italic_I start_POSTSUBSCRIPT roman_HH end_POSTSUBSCRIPT end_ARG end_CELL end_ROW (3)

Second Harmonic Generation Measurements

Second harmonic generation (SHG) measurements were performed in a transmission geometry using the home-built setup, as shown in Supplementary Information Section 12. An idler beam generated from a TOPAS Prime Automated OPA (Spectra-Physics, USA) with a 5 kHz repetition rate, a 100 fs pulse duration, and a tunable wavelength from 1760 nm–2100 nm was used. A BBO nonlinear optical crystal was used to convert the laser wavelength to 880 nm–1050 nm via second harmonic generation, followed by a 1326 nm short pass optical filter to block the fundamental laser beam. The laser then passed through a Berek polarization compensator (Model 5540, Newport Corp.) set in half-wave plate mode to make sure the laser polarization was corrected to be in the yz𝑦𝑧yzitalic_y italic_z (yzsuperscript𝑦superscript𝑧y^{\prime}z^{\prime}italic_y start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT) plane in the laboratory (sample) coordinate system, as indicated in Fig. 2a. A convex lens with a focal length of 75 mm was utilized to focus the laser beam onto the sample. The diameter of the laser spot was measured to be around 70 μμ\upmuroman_μm, varying with the laser wavelength, giving a probing area on the order of 103μμ\upmuroman_μm2. The alignment of the sample was also kept in the yz𝑦𝑧yzitalic_y italic_z plane and tilted along the x𝑥xitalic_x-axis in the laboratory coordinate system so that chirality-induced SHG can be generated. The SHG signal was extracted by a 785 nm short-pass filter and a 633 nm short-pass filter. A polarizer was put right before the CMOS camera detector (PM001, Brigates Microelectronics, China) to analyze the signal’s polarization. The detector recorded the SHG signal and by integrating the count from each pixel, subtracting the background, and correcting the quantum efficiency of the detector, the actual SHG photon count can be acquired.

\bmhead

Data Availability Source data for the figures in the main text are provided with the paper. All other data supporting the plots within this paper and other study findings are available from the corresponding author upon reasonable request.

\bmhead

Acknowledgements R.X., J.L., and H.Z. acknowledge the U.S. National Science Foundation (DMR-2240106) and Robert A. Welch Foundation (C-2128). J.D., V.L., A.B., F.T., V.P., and J.K. acknowledge the U.S. National Science Foundation (PIRE-2230727). J.D., A.B., F.T., and J.K. also acknowledge support from the Robert A. Welch Foundation (C-1509), the Air Force Office of Scientific Research (FA9550-22-1-0382), and the Chan Zuckerberg Initiative (WU-21-357). Y.Y. and K.Y. acknowledge support from JSPS KAKENHI (JP20H02573, JP21H05017, JP22H05469 and JP23H00259), JST CREST (JPMJCR17I5), and US-JAPAN PIRE collaboration(JPJSJRP20221202). W.T. and J.L. acknowledge Robert A. Welch Foundation (C-1716). EB is supported by the Graduate Research Fellowship Program (GRFP) from U.S. National Science Foundation under grant number DGE-1842494. R.S. acknowledges support from JSPS KAKENHI (JP22H00283) and Yushan Fellow Program from Taiwan.

\bmhead

Author Contributions A.B., J.K., and H.Z. conceived the project. Y.Y. and K.Y. synthesized the enantiomer-pure (6,5)- CNT suspension. J.D. fabricated the CNT film samples. K.Y., Y.Y., and J.D. performed circular dichroism spectroscopy. J.D. and N.H. measured the refractive index of the enantiomer-pure aligned CNT thin film. R.X., J.D., and N.H. measured the polarized attenuation spectra. J.D. and R.X. performed AFM measurements. J.D., W.T., and R.X. performed SEM measurements. J.D. performed polarized Raman measurements. R.X. performed SHG experiments and analyzed the data. V.L. conducted the many-body atomistic calculations under the supervision of V.P. R.X., J.D., A.S., M.T.,E.B., J.L.,F.T., V.P., A.B., and H.Z. wrote the manuscript with input from all authors. All authors discussed the data.

\bmhead

Competing Interests The authors declare no competing financial interests.

References

  • \bibcommenthead
  • (1) Circular Dichroism: Principles and Applications, 2nd Edition | Wiley
  • (2) Ahn, J., Lee, E., Tan, J., Yang, W., Kim, B., Moon, J.: A new class of chiral semiconductors: chiral-organic-molecule-incorporating organic–inorganic hybrid perovskites. Materials Horizons 4(5), 851–856 (2017)
  • (3) Rikken, G.L.J.A., Raupach, E.: Observation of magneto-chiral dichroism. Nature 390(6659), 493–494 (1997)
  • (4) Train, C., Gheorghe, R., Krstic, V., Chamoreau, L.-M., Ovanesyan, N.S., Rikken, G.L.J.A., Gruselle, M., Verdaguer, M.: Strong magneto-chiral dichroism in enantiopure chiral ferromagnets. Nature Materials 7(9), 729–734 (2008)
  • (5) Naaman, R., Waldeck, D.H.: Chiral-Induced Spin Selectivity Effect. The Journal of Physical Chemistry Letters 3(16), 2178–2187 (2012)
  • (6) Yang, C., Li, Y., Zhou, S., Guo, Y., Jia, C., Liu, Z., Houk, K.N., Dubi, Y., Guo, X.: Real-time monitoring of reaction stereochemistry through single-molecule observations of chirality-induced spin selectivity. Nature Chemistry, 1–8 (2023)
  • (7) Yang, Y., da Costa, R.C., Fuchter, M.J., Campbell, A.J.: Circularly polarized light detection by a chiral organic semiconductor transistor. Nature Photonics 7(8), 634–638 (2013)
  • (8) Brandt, J.R., Salerno, F., Fuchter, M.J.: The added value of small-molecule chirality in technological applications. Nature Reviews Chemistry 1(6), 1–12 (2017)
  • (9) Yang, S.-H., Naaman, R., Paltiel, Y., Parkin, S.S.P.: Chiral spintronics. Nature Reviews Physics 3(5), 328–343 (2021)
  • (10) Wei, Q., Ning, Z.: Chiral Perovskite Spin-Optoelectronics and Spintronics: Toward Judicious Design and Application. ACS Materials Letters 3(9), 1266–1275 (2021)
  • (11) Singh, S., Bonner, W.A., Van Uitert, L.G.: Violation of Kleinman’s symmetry condition in paratellurite. Physics Letters A 38(6), 407–408 (1972)
  • (12) Chemla, D.S., Jerphagnon, J.: Optical Second‐Harmonic Generation in Paratellurite and Kleinman’s Symmetry Relations. Applied Physics Letters 20(6), 222–223 (2003)
  • (13) Yuan, C., Li, X., Semin, S., Feng, Y., Rasing, T., Xu, J.: Chiral Lead Halide Perovskite Nanowires for Second-Order Nonlinear Optics. Nano Letters 18(9), 5411–5417 (2018)
  • (14) Yao, L., Zeng, Z., Cai, C., Xu, P., Gu, H., Gao, L., Han, J., Zhang, X., Wang, X., Wang, X., Pan, A., Wang, J., Liang, W., Liu, S., Chen, C., Tang, J.: Strong Second- and Third-Harmonic Generation in 1D Chiral Hybrid Bismuth Halides. Journal of the American Chemical Society 143(39), 16095–16104 (2021)
  • (15) Ge, F., Li, B.-H., Cheng, P., Li, G., Ren, Z., Xu, J., Bu, X.-H.: Chiral Hybrid Copper(I) Halides for High Efficiency Second Harmonic Generation with a Broadband Transparency Window. Angewandte Chemie 134(10), 202115024 (2022)
  • (16) Boyd, R.W.: Nonlinear Optics. Academic Press, Boston (1992)
  • (17) Cheng, M., Wu, S., Zhu, Z.-Z., Guo, G.-Y.: Large second-harmonic generation and linear electro-optic effect in trigonal selenium and tellurium. Physical Review B 100(3), 035202 (2019)
  • (18) Londoño-Calderon, A., Williams, D.J., Schneider, M.M., Savitzky, B.H., Ophus, C., Ma, S., Zhu, H., Pettes, M.T.: Intrinsic helical twist and chirality in ultrathin tellurium nanowires. Nanoscale 13(21), 9606–9614 (2021)
  • (19) Wang, H., Mao, Y., Kislyakov, I.M., Dong, N., Chen, C., Wang, J.: Anisotropic Raman scattering and intense broadband second-harmonic generation in tellurium nanosheets. Optics Letters 46(8), 1812–1815 (2021)
  • (20) Sherman, G., Coleman, P.: Absolute measurement of the second-harmonic generation nonlinear susceptibility of tellurium at 28.028.028.028.0 mm 9(3), 403–409
  • (21) Avouris, P., Freitag, M., Perebeinos, V.: Carbon-nanotube photonics and optoelectronics. Nature Photon 2(6), 341–350 (2008)
  • (22) Blackburn, J.L., Ferguson, A.J., Cho, C., Grunlan, J.C.: Carbon-Nanotube-Based Thermoelectric Materials and Devices. Advanced Materials 30(11), 1704386 (2018)
  • (23) Li, Y.: Carbon Nanotube Research in Its 30th Year. ACS Nano 15(6), 9197–9200 (2021)
  • (24) Baydin, A., Tay, F., Fan, J., Manjappa, M., Gao, W., Kono, J.: Carbon Nanotube Devices for Quantum Technology. Materials, 26 (2022)
  • (25) Guo, G.Y., Chu, K.C., Wang, D.-s., Duan, C.-g.: Linear and nonlinear optical properties of carbon nanotubes from first-principles calculations. Physical Review B 69(20), 205416 (2004)
  • (26) Rérat, M., Karamanis, P., Civalleri, B., Maschio, L., Lacivita, V., Kirtman, B.: Ab Initio Calculation of Nonlinear Optical Properties for Chiral Carbon Nanotubes. Second Harmonic Generation and Dc-Pockels Effect. Theor Chem Acc 137(2), 17 (2018)
  • (27) Weisman, R.B., Kono, J.: Introduction to Optical Spectroscopy of Single-Wall Carbon Nanotubes. In: Handbook of Carbon Nanomaterials. World Scientific Series on Carbon Nanoscience, vol. Volume 9 & 10, pp. 1–43. World Scientific, Singapore (2017)
  • (28) Nishihara, T., Takakura, A., Shimasaki, M., Matsuda, K., Tanaka, T., Kataura, H., Miyauchi, Y.: Empirical formulation of broadband complex refractive index spectra of single-chirality carbon nanotube assembly. Nanophotonics 11(5), 1011–1020 (2022)
  • (29) Huttunen, M.J., Herranen, O., Johansson, A., Jiang, H., Mudimela, P.R., Myllyperkiö, P., Bautista, G., Nasibulin, A.G., Kauppinen, E.I., Ahlskog, M., Kauranen, M., Pettersson, M.: Measurement of Optical Second-Harmonic Generation from an Individual Single-Walled Carbon Nanotube. New J. Phys. 15(8), 083043 (2013)
  • (30) Okawara, S., Kudo, M., Miyamori, M., Ohno, S., Shimazu, Y., Tanaka, M., Suzuki, T.: Second-Harmonic Generation from Supported Carbon Nanotube Films Grown by Chemical Vapor Deposition on Fused Silica. Jpn. J. Appl. Phys. 58(3), 032006 (2019)
  • (31) Dominicis, L.D., Fantoni, R., Botti, S., Asilyan, L.S., Ciardi, R., Fiori, A., Appolloni, R.: Symmetry point group description of second harmonic generation in carbon nanotubes. Laser Physics Letters 1(4), 172 (2004)
  • (32) Su, H.M., Ye, J.T., Tang, Z.K., Wong, K.S.: Resonant Second-Harmonic Generation in Monosized and Aligned Single-Walled Carbon Nanotubes. Phys. Rev. B 77(12), 125428 (2008)
  • (33) Murakami, Y., Kono, J.: Nonlinear Photoluminescence Excitation Spectroscopy of Carbon Nanotubes: Exploring the Upper Density Limit of One-Dimensional Excitons. Physical Review Letters 102(3), 037401 (2009)
  • (34) Janisch, C., Wang, Y., Ma, D., Mehta, N., Elías, A.L., Perea-López, N., Terrones, M., Crespi, V., Liu, Z.: Extraordinary Second Harmonic Generation in Tungsten Disulfide Monolayers. Scientific Reports 4(1), 5530 (2014)
  • (35) Kumar, N., Najmaei, S., Cui, Q., Ceballos, F., Ajayan, P.M., Lou, J., Zhao, H.: Second harmonic microscopy of monolayer MoS2. Physical Review B 87(16), 161403 (2013)
  • (36) You, J.W., Bongu, S.R., Bao, Q., Panoiu, N.C.: Nonlinear optical properties and applications of 2D materials: theoretical and experimental aspects. Nanophotonics 8(1), 63–97 (2019)
  • (37) Kim, B., Jin, J., Wang, Z., He, L., Christensen, T., Mele, E.J., Zhen, B.: Three-dimensional nonlinear optical materials from twisted two-dimensional van der Waals interfaces. Nature Photonics 18(1), 91–98 (2024)
  • (38) Wu, L., Patankar, S., Morimoto, T., Nair, N.L., Thewalt, E., Little, A., Analytis, J.G., Moore, J.E., Orenstein, J.: Giant anisotropic nonlinear optical response in transition metal monopnictide Weyl semimetals. Nature Physics 13(4), 350–355 (2017)
  • (39) Lodahl, P., Mahmoodian, S., Stobbe, S., Rauschenbeutel, A., Schneeweiss, P., Volz, J., Pichler, H., Zoller, P.: Chiral quantum optics. Nature 541(7638), 473–480 (2017)
  • (40) Yomogida, Y., Tanaka, T., Zhang, M., Yudasaka, M., Wei, X., Kataura, H.: Industrial-scale separation of high-purity single-chirality single-wall carbon nanotubes for biological imaging. Nature Communications 7(1), 12056 (2016)
  • (41) Yomogida, Y., Tanaka, T., Tsuzuki, M., Wei, X., Kataura, H.: Automatic Sorting of Single-Chirality Single-Wall Carbon Nanotubes Using Hydrophobic Cholates: Implications for Multicolor Near-Infrared Optical Technologies. ACS Applied Nano Materials 3(11), 11289–11297 (2020)
  • (42) Wei, X., Tanaka, T., Hirakawa, T., Tsuzuki, M., Wang, G., Yomogida, Y., Hirano, A., Kataura, H.: High-yield and high-throughput single-chirality enantiomer separation of single-wall carbon nanotubes. Carbon 132, 1–7 (2018)
  • (43) Weisman, R.B., Kono, J.: Optical Properties of Carbon Nanotubes: A Volume Dedicated to the Memory of Professor Mildred Dresselhaus. World Scientific, Singapore (2019)
  • (44) Wei, X., Tanaka, T., Hirakawa, T., Yomogida, Y., Kataura, H.: Determination of Enantiomeric Purity of Single-Wall Carbon Nanotubes Using Flavin Mononucleotide. Journal of the American Chemical Society 139(45), 16068–16071 (2017)
  • (45) Wei, X., Luo, X., Li, S., Zhou, W., Xie, S., Liu, H.: Length-Dependent Enantioselectivity of Carbon Nanotubes by Gel Chromatography. ACS Nano 17(9), 8393–8402 (2023)
  • (46) He, X., Gao, W., Xie, L., Li, B., Zhang, Q., Lei, S., Robinson, J.M., Hároz, E.H., Doorn, S.K., Wang, W., Vajtai, R., Ajayan, P.M., Adams, W.W., Hauge, R.H., Kono, J.: Wafer-scale monodomain films of spontaneously aligned single-walled carbon nanotubes. Nature Nanotechnology 11(7), 633–638 (2016)
  • (47) Maultzsch, J., Telg, H., Reich, S., Thomsen, C.: Radial breathing mode of single-walled carbon nanotubes: Optical transition energies and chiral-index assignment. Phys. Rev. B 72, 205438 (2005)
  • (48) Katsutani, F., Gao, W., Li, X., Ichinose, Y., Yomogida, Y., Yanagi, K., Kono, J.: Direct observation of cross-polarized excitons in aligned single-chirality single-wall carbon nanotubes. Physical Review B 99(3), 035426 (2019)
  • (49) Qian, H., Li, S., Chen, C.-F., Hsu, S.-W., Bopp, S.E., Ma, Q., Tao, A.R., Liu, Z.: Large optical nonlinearity enabled by coupled metallic quantum wells. Light: Science & Applications 8(1), 13 (2019)
  • (50) Fu, Q., Cong, X., Xu, X., Zhu, S., Zhao, X., Liu, S., Yao, B., Xu, M., Deng, Y., Zhu, C., Wang, X., Kang, L., Zeng, Q., Lin, M.-L., Wang, X., Tang, B., Yang, J., Dong, Z., Liu, F., Xiong, Q., Zhou, J., Wang, Q., Li, X., Tan, P.-H., Tay, B.K., Liu, Z.: Berry Curvature Dipole Induced Giant Mid-Infrared Second-Harmonic Generation in 2D Weyl Semiconductor. Advanced Materials 35(46), 2306330 (2023)
  • (51) Zhang, Y., Gao, B., Lepage, D., Tong, Y., Wang, P., Xia, W., Niu, J., Feng, Y., Chen, H., Qian, H.: Large second-order susceptibility from a quantized indium tin oxide monolayer. Nature Nanotechnology, 1–8 (2024)
  • (52) Hussain, S.A.: Comparison of Graphene and Carbon Nanotube Saturable Absorbers for Wavelength and Pulse Duration Tunability. Scientific Reports 9(1), 17282 (2019)
  • (53) Wang, J., Liang, X., Hu, G., Zheng, Z., Lin, S., Ouyang, D., Wu, X., Yan, P., Ruan, S., Sun, Z., Hasan, T.: 152 fs nanotube-mode-locked thulium-doped all-fiber laser. Scientific Reports 6(1), 28885 (2016)
  • (54) Malapanis, A., Perebeinos, V., Sinha, D.P., Comfort, E., Lee, J.U.: Quantum efficiency and capture cross section of first and second excitonic transitions of single-walled carbon nanotubes measured through photoconductivity. Nano Letters 13, 3531–3538 (2013)
  • (55) Rohlfing, M., Louie, S.G.: Electron-hole excitations and optical spectra from first principles. Phys. Rev. B 62, 4927–4944 (2000)
  • (56) Hertel, T., Perebeinos, V., Crochet, J., Arnold, K., Kappes, M., Avouris, P.: Intersubband decay of 1-d exciton resonances in carbon nanotubes. Nano Letters 8, 87–91 (2008)
  • (57) Boyd, R.W. (ed.): Nonlinear Optics (Fourth Edition). Academic Press, London (2020)
  • (58) Pedersen, T.G., Pedersen, K.: Systematic tight-binding study of optical second-harmonic generation in carbon nanotubes. Physical Review B 79(3), 035422 (2009)
  • (59) Guo, Q., Qi, X.-Z., Zhang, L., Gao, M., Hu, S., Zhou, W., Zang, W., Zhao, X., Wang, J., Yan, B., Xu, M., Wu, Y.-K., Eda, G., Xiao, Z., Yang, S.A., Gou, H., Feng, Y.P., Guo, G.-C., Zhou, W., Ren, X.-F., Qiu, C.-W., Pennycook, S.J., Wee, A.T.S.: Ultrathin quantum light source with van der Waals NbOCl2 crystal. Nature 613(7942), 53–59 (2023). https://doi.org/10.1038/s41586-022-05393-7
  • (60) Chernikov, A., van der Zande, A.M., Hill, H.M., Rigosi, A.F., Velauthapillai, A., Hone, J., Heinz, T.F.: Electrical Tuning of Exciton Binding Energies in Monolayer WS2. Physical Review Letters 115(12), 126802 (2015)
  • (61) Lin, Y., Ling, X., Yu, L., Huang, S., Hsu, A.L., Lee, Y.-H., Kong, J., Dresselhaus, M.S., Palacios, T.: Dielectric Screening of Excitons and Trions in Single-Layer MoS2. Nano Letters 14(10), 5569–5576 (2014)
  • (62) Ghosh, S., Bachilo, S.M., Weisman, R.B.: Advanced sorting of single-walled carbon nanotubes by nonlinear density-gradient ultracentrifugation. Nature Nanotechnology 5(6), 443–450 (2010)
  • (63) Ao, G., Streit, J.K., Fagan, J.A., Zheng, M.: Differentiating Left- and Right-Handed Carbon Nanotubes by DNA. Journal of the American Chemical Society 138(51), 16677–16685 (2016). Publisher: American Chemical Society
  • (64) Samsonidze, G.G., Grüneis, A., Saito, R., Jorio, A., Souza Filho, A.G., Dresselhaus, G., Dresselhaus, M.S.: Interband optical transitions in left- and right-handed single-wall carbon nanotubes. Phys. Rev. B 69, 205402 (2004)
  • (65) Zamora-Ledezma, C., Blanc, C., Maugey, M., Zakri, C., Poulin, P., Anglaret, E.: Anisotropic Thin Films of Single-Wall Carbon Nanotubes from Aligned Lyotropic Nematic Suspensions. Nano Letters 8(12), 4103–4107 (2008)